ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4246
Committed: Thu Dec 11 01:24:49 2014 UTC (9 years, 6 months ago) by plouden
Content type: application/x-tex
File size: 44787 byte(s)
Log Message:
Fixed changes from today, still need to work on SPC/E justification.

File Contents

# Content
1 %% PNAStwoS.tex
2 %% Sample file to use for PNAS articles prepared in LaTeX
3 %% For two column PNAS articles
4 %% Version1: Apr 15, 2008
5 %% Version2: Oct 04, 2013
6
7 %% BASIC CLASS FILE
8 \documentclass{pnastwo}
9
10 %% ADDITIONAL OPTIONAL STYLE FILES Font specification
11
12 %\usepackage{PNASTWOF}
13 \usepackage[version=3]{mhchem}
14 \usepackage[round,numbers,sort&compress]{natbib}
15 \usepackage{fixltx2e}
16 \bibpunct{(}{)}{,}{n}{,}{,}
17 \bibliographystyle{pnas2011}
18
19 %% OPTIONAL MACRO DEFINITIONS
20 \def\s{\sigma}
21 %%%%%%%%%%%%
22 %% For PNAS Only:
23 %\url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
24 \copyrightyear{2014}
25 \issuedate{Issue Date}
26 \volume{Volume}
27 \issuenumber{Issue Number}
28 %\setcounter{page}{2687} %Set page number here if desired
29 %%%%%%%%%%%%
30
31 \begin{document}
32
33 \title{Friction at water / ice-I\textsubscript{h} interfaces: Do the
34 different facets of ice have different hydrophilicities?}
35
36 \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
37 IN 46556}
38 \and
39 J. Daniel Gezelter\affil{1}{}}
40
41 \contributor{Submitted to Proceedings of the National Academy of Sciences
42 of the United States of America}
43
44 %%%Newly updated.
45 %%% If significance statement need, then can use the below command otherwise just delete it.
46 %\significancetext{RJSM and ACAC developed the concept of the study. RJSM conducted the analysis, data interpretation and drafted the manuscript. AGB contributed to the development of the statistical methods, data interpretation and drafting of the manuscript.}
47
48 \maketitle
49
50 \begin{article}
51 \begin{abstract}
52 In this paper we present evidence that some of the crystal facets
53 of ice-I$_\mathrm{h}$ posess structural features that can halve
54 the effective hydrophilicity of the ice/water interface. The
55 spreading dynamics of liquid water droplets on ice facets exhibits
56 long-time behavior that differs substantially for the prismatic
57 $\{1~0~\bar{1}~0\}$ and secondary prism $\{1~1~\bar{2}~0\}$ facets
58 when compared with the basal $\{0001\}$ and pyramidal
59 $\{2~0~\bar{2}~1\}$ facets. We also present the results of
60 simulations of solid-liquid friction of the same four crystal
61 facets being drawn through liquid water. Both simulation
62 techniques provide evidence that the two prismatic faces have an
63 effective surface area in contact with the liquid water of
64 approximately half of the total surface area of the crystal. The
65 ice / water interfacial widths for all four crystal facets are
66 similar (using both structural and dynamic measures), and were
67 found to be independent of the shear rate. Additionally,
68 decomposition of orientational time correlation functions show
69 position-dependence for the short- and longer-time decay
70 components close to the interface.
71 \end{abstract}
72
73 \keywords{ice | water | interfaces | hydrophobicity}
74 \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
75 reverse non-equilibrium molecular dynamics}
76
77 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
78 based on the strength of the interactions with water. Hydrophobic
79 surfaces do not have strong enough interactions with water to overcome
80 the internal attraction between molecules in the liquid phase, and the
81 degree of hydrophilicity of a surface can be described by the extent a
82 droplet can spread out over the surface. The contact angle formed
83 between the solid and the liquid depends on the free energies of the
84 three interfaces involved, and is given by Young's
85 equation.\cite{Young05}
86 \begin{equation}\label{young}
87 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
88 \end{equation}
89 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
90 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces
91 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
92 to hydrophobic surfaces with low wettability, while small contact
93 angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
94 Experimentally, measurements of the contact angle of sessile drops is
95 often used to quantify the extent of wetting on surfaces with
96 thermally selective wetting
97 characteristics.\cite{Tadanaga00,Liu04,Sun04}
98
99 Nanometer-scale structural features of a solid surface can influence
100 the hydrophilicity to a surprising degree. Small changes in the
101 heights and widths of nano-pillars can change a surface from
102 superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
103 \sim 0^{\circ}$.\cite{Koishi09} This is often referred to as the
104 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
105 electrically tunable Cassie-Baxter and Wenzel states have also been
106 observed.\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}
107 Luzar and coworkers have modeled these transitions on nano-patterned
108 surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
109 change in contact angle is due to the field-induced perturbation of
110 hydrogen bonding at the liquid/vapor interface.\cite{Daub07}
111
112 One would expect the interfaces of ice to be highly hydrophilic (and
113 possibly the most hydrophilic of all solid surfaces). In this paper we
114 present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
115 have structural features that can halve the effective hydrophilicity.
116 Our evidence for this comes from molecular dynamics (MD) simulations
117 of the spreading dynamics of liquid droplets on these facets, as well
118 as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
119 solid-liquid friction.
120
121 Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
122 extensively using computer simulations. Haymet \textit{et al.}
123 characterized and measured the width of these interfaces for the
124 SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
125 CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
126 both neat water and with solvated
127 ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
128 studied the width of basal/water and prismatic/water
129 interfaces~\cite{Nada95} as well as crystal restructuring at
130 temperatures approaching the melting point~\cite{Nada00}.
131
132 The surface of ice exhibits a premelting layer, often called a
133 quasi-liquid layer (QLL), at temperatures near the melting point. MD
134 simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
135 found QLL widths of approximately 10 \AA\ at 3 K below the melting
136 point.\cite{Conde08} Similarly, Limmer and Chandler have used the mW
137 water model~\cite{Molinero09} and statistical field theory to estimate
138 QLL widths at similar temperatures to be about 3 nm.\cite{Limmer14}
139
140 Recently, Sazaki and Furukawa have developed a technique using laser
141 confocal microscopy combined with differential interference contrast
142 microscopy that has sufficient spatial and temporal resolution to
143 visulaize and quantitatively analyze QLLs on ice crystals at
144 temperatures near melting.\cite{Sazaki10} They have found the width of
145 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
146 wide. They have also seen the formation of two immiscible QLLs, which
147 displayed different dynamics on the crystal surface.\cite{Sazaki12}
148
149 There is now significant interest in the \textit{tribological}
150 properties of ice/ice and ice/water interfaces in the geophysics
151 community. Understanding the dynamics of solid-solid shearing that is
152 mediated by a liquid layer\cite{Cuffey99, Bell08} will aid in
153 understanding the macroscopic motion of large ice
154 masses.\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}
155
156 Using molecular dynamics simulations, Samadashvili has recently shown
157 that when two smooth ice slabs slide past one another, a stable
158 liquid-like layer develops between them.\cite{Samadashvili13} In a
159 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
160 through liquid water have provided quantitative estimates of the
161 solid-liquid kinetic friction coefficients.\cite{Louden13} These
162 displayed a factor of two difference between the basal and prismatic
163 facets. The friction was found to be independent of shear direction
164 relative to the surface orientation. We attributed facet-based
165 difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
166 prismatic face which reduces the effective surface area of the ice
167 that is in direct contact with liquid water.
168
169 In the sections that follow, we outline the methodology used to
170 simulate droplet-spreading dynamics using standard MD and tribological
171 properties using RNEMD simulations. These simulation methods give
172 complementary results that point to the prismatic and secondary prism
173 facets having roughly half of their surface area in direct contact
174 with the liquid.
175
176 \section{Methodology}
177 \subsection{Construction of the Ice / Water Interfaces}
178 To construct the four interfacial ice/water systems, a proton-ordered,
179 zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
180 H-atoms and lone pairs was constructed using Structure 6 of Hirsch and
181 Ojam\"{a}e's set of orthorhombic representations for
182 ice-I$_{h}$.\cite{Hirsch04} This crystal structure was cleaved along
183 the four different facets being studied. The exposed face was
184 reoriented normal to the $z$-axis of the simulation cell, and the
185 structures were and extended to form large exposed facets in
186 rectangular box geometries. Liquid water boxes were created with
187 identical dimensions (in $x$ and $y$) as the ice, and a $z$ dimension
188 of three times that of the ice block, and a density corresponding to
189 $\sim$ 1 g / cm$^3$. Each of the ice slabs and water boxes were
190 independently equilibrated, and the resulting systems were merged by
191 carving out any liquid water molecules within 3 \AA\ of any atoms in
192 the ice slabs. Each of the combined ice/water systems were then
193 equilibrated at 225K, which is the liquid-ice coexistence temperature
194 for SPC/E water.\cite{Bryk02} Ref. \citealp{Louden13} contains a more
195 detailed explanation of the construction of ice/water interfaces. The
196 resulting dimensions, number of ice, and liquid water molecules
197 contained in each of these systems can be seen in Table
198 \ref{tab:method}.
199
200 Mostly avoids spurious crystalline morphologies like ice-i and ice-B.
201
202 The SPC/E water model\cite{Berendsen87} has been extensively
203 characterized over a wide range of liquid
204 conditions,\cite{Arbuckle02, Kuang12} and its phase diagram has been well studied.\cite{Baez95,Bryk04b,Sanz04b}
205 The free energies \cite{Baez95} and melting points
206 \cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fernandez06,Abascal07,Vrbka07}
207 of various crystal structures have also been calculated.
208 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water
209 interface using the SPC/E water model, and has seen good agreement for
210 structural and dynamic measurements of the interfacial width when compared with more
211 expensive water models. For these reasons, the SPC/E water
212 model was used in this study.
213
214 \subsection{Shearing simulations (interfaces in bulk water)}
215 % Should we mention number of runs, sim times, etc. ?
216 To perform the shearing simulations, the velocity shearing and scaling
217 varient of reverse nonequilibrium molecular dynamics (VSS-RNEMD) was
218 conducted. This method performs a series of simultaneous velocity
219 exchanges between two regions of the simulation cell, to
220 simultaneously create a velocity and temperature gradient. The thermal
221 gradient is necessary when performing shearing simulations as to
222 prevent frictional heating from the shear from melting the
223 interface. For more details on the VSS-RNEMD method please refer to a
224 pervious paper\cite{Louden13}.
225
226 The computational details performed here were equivalent to those reported
227 in a previous publication\cite{Louden13}, with the following changes.
228 VSS-RNEMD moves were attempted every 2 fs instead of every 50 fs. This was done to minimize
229 the magnitude of each individual VSS-RNEMD perturbation to the system.
230 All pyramidal simulations were performed under the canonical (NVT) ensamble
231 except those during which configurations were accumulated for the orientational correlation
232 function, which were performed under the microcanonical (NVE) ensamble. All
233 secondary prismatic simulations were performed under the NVE ensamble.
234
235 \subsection{Droplet simulations}
236 Ice interfaces with a thickness of $\sim 20 \AA$ were created as
237 described above, but were not solvated in a liquid box. The crystals
238 were then replicated along the $x$ and $y$ axes (parallel to the
239 surface) until a large surface had been created. The sizes and
240 numbers of molecules in each of the surfaces is given in Table
241 \ref{tab:ice_sheets}. Weak translational restraining potentials with
242 spring constants of 1.5 to 4.0 UNITS were applied to the center of mass of each
243 molecule in order to prevent surface melting, although the molecules
244 were allowed to reorient freely. A water doplet containing 2048 SPC/E
245 molecules was created separately. Droplets of this size can produce
246 agreement with the Young contact angle extrapolated to an infinite
247 drop size\cite{Daub10}. The surfaces and droplet were independently
248 equilibrated to 225 K, at which time the droplet was placed 3-5 \AA\
249 above the surface. Five statistically independent simulations were
250 carried out for each facet, and the droplet was placed at unique $x$
251 and $y$ locations for each of these simulations. Each simulation was
252 5 ns in length and conducted in the microcanonical (NVE) ensemble.
253
254 \section{Results and discussion}
255 \subsection{Dynamic water contact angle}
256
257 To determine the extent of wetting for each of the four crystal
258 facets, water contact angle simuations were performed. Contact angles
259 were obtained from these simulations by two methods. In the first
260 method, the contact angle was obtained from the $z$-center of mass
261 ($z_{c.m.}$) of the water droplet as described by Hautman and Klein\cite{Hautman91}
262 and utilized by Hirvi and Pakkanen in their investigation of water
263 droplets on polyethylene and poly(vinyl chloride)
264 surface\cite{Hirvi06}. At each snapshot of the simulation, the contact
265 angle, $\theta$, was found by
266 \begin{equation}\label{contact_1}
267 \langle z_{c.m.}\rangle = 2^{-4/3}R_{0}\bigg(\frac{1-cos\theta}{2+cos\theta}\bigg)^{1/3}\frac{3+cos\theta}{2+cos\theta} ,
268 \end{equation}
269 where $R_{0}$ is the radius of the free water droplet. In the second
270 method, the contact angle was obtained from fitting the droplet's
271 $z$-profile after radial averaging to a
272 circle as described by Ruijter, Blake, and
273 Coninck\cite{Ruijter99}. At each snapshot of the simulation the water droplet was
274 broken into bins, and the location of bin containing half-bulk density was
275 stored. Due to fluctuations close to the ice, all bins located within
276 2.0 \AA\ of the ice were discarded. The remaining stored bins were
277 then fit by a circle, whose tangential intersection with the ice plane could
278 be used to calculate the water
279 contact angle. These results proved noisey and unreliable when
280 compared with the first method, for these purposes we omit the data
281 from the second method.
282
283 The resulting water contact angle profiles generated by the first method
284 had an initial value of 180$^{o}$, and decayed over time. Each of
285 these profiles were fit to a biexponential decay, with a short time
286 piece to account for the water droplet initially adhering to the
287 surface, a long time piece describing the spreading of the droplet
288 over the surface, and an added constant to capture the infinite
289 decay of the contact angle. We have found that the rate of the water
290 droplet spreading across all four crystal facets to be $\approx$ 0.7
291 ns$^{-1}$. However, the basal and pyramidal facets
292 had an infinite decay value for $\theta$ of $\approx$ 35$^{o}$, while
293 prismatic and secondary prismatic had values for $\theta$ near
294 43$^{o}$ as seen in Table \ref{tab:kappa}. These results indicate that the
295 basal and pyramidal facets are more hydrophilic than the prismatic and
296 secondary prismatic, and surprisingly, that the differential hydrophilicities of
297 the crystal facets is not reflected in the spreading rate of the droplet.
298 % This is in good agreement with our calculations of friction
299 % coefficients, in which the basal
300 % and pyramidal had a higher coefficient of kinetic friction than the
301 % prismatic and secondary prismatic. Due to this, we beleive that the
302 % differences in friction coefficients can be attributed to the varying
303 % hydrophilicities of the facets.
304
305 \subsection{Coefficient of friction of the interfaces}
306 While investigating the kinetic coefficient of friction, there was found
307 to be a dependence for $\mu_k$
308 on the temperature of the liquid water in the system. We believe this
309 dependence
310 arrises from the sharp discontinuity of the viscosity for the SPC/E model
311 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
312 a weighting to the interfacial friction coefficient, $\kappa$ by the
313 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
314 relates the shear stress with the relative velocity of the fluid normal to the
315 interface:
316 \begin{equation}\label{Shenyu-13}
317 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
318 \end{equation}
319 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
320 in the
321 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
322 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
323 fluid can be determined under a linear response of the momentum
324 gradient to the applied shear stress by
325 \begin{equation}\label{Shenyu-11}
326 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
327 \end{equation}
328 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
329 expression for $\kappa$,
330 \begin{equation}\label{kappa-1}
331 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
332 \end{equation}
333 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
334 giving us
335 \begin{equation}\label{kappa-2}
336 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
337 \end{equation}
338
339 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
340 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
341 equilibrated to 225K,
342 and 5 unique shearing experiments were performed. Each experiment was
343 conducted in the NVE and were 5 ns in
344 length. The VSS were attempted every timestep, which was set to 2 fs.
345 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
346 roughly ten times larger than the value found for 280 K SPC/E bulk water by
347 Kuang\cite{Kuang12}.
348
349 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
350 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
351 is an indication of how 'slippery' the interface is.
352 \begin{equation}\label{kappa-3}
353 \kappa = \frac{\eta}{\delta}
354 \end{equation}
355 In each of the systems, the interfacial temperature was kept fixed to 225K,
356 which ensured the viscosity of the fluid at the
357 interace was approximately the same. Thus, any significant variation in
358 $\kappa$ between
359 the systems indicates differences in the 'slipperiness' of the interfaces.
360 As each of the ice systems are sheared relative to liquid water, the
361 'slipperiness' of the interface can be taken as an indication of how
362 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
363 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
364 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
365 similar values of $\kappa \approx$ 0.0006
366 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
367 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
368 were found for the prismatic and secondary prismatic systems.
369 These results indicate that the basal and pyramidal facets are
370 more hydrophilic than the prismatic and secondary prismatic facets.
371
372 \subsection{Interfacial width}
373 In the literature there is good agreement that between the solid ice and
374 the bulk water, there exists a region of 'slush-like' water molecules.
375 In this region, the water molecules are structurely distinguishable and
376 behave differently than those of the solid ice or the bulk water.
377 The characteristics of this region have been defined by both structural
378 and dynamic properties; and its width has been measured by the change of these
379 properties from their bulk liquid values to those of the solid ice.
380 Examples of these properties include the density, the diffusion constant, and
381 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
382
383 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
384 perturb the momenta of the water molecules in
385 the systems, parameters that depend on translational motion may give
386 faulty results. A stuructural parameter will be less effected by the
387 VSS-RNEMD perturbations to the system. Due to this, we have used the
388 local tetrahedral order parameter (Eq 5\cite{Louden13}) to quantify the width of the interface,
389 which was originally described by Kumar\cite{Kumar09} and
390 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
391 of ice/water interfaces.\cite{Bryk04b}
392
393 To determine the width of the interfaces, each of the systems were
394 divided into 100 artificial bins along the
395 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
396 time-averaged for each of the bins, resulting in a tetrahedrality profile of
397 the system. These profiles are shown across the $z$-dimension of the systems
398 in panel $a$ of Figures \ref{fig:pyrComic}
399 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
400 (0,1), where a larger value indicates a more tetrahedral environment.
401 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
402 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
403 fit using a hyperbolic tangent (Eq. 6\cite{Louden13}) designed to smoothly fit the
404 bulk to ice
405 transition, while accounting for the thermal influence on the profile by the
406 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
407 resulting thermal and velocity gradients from an imposed kinetic energy and
408 momentum fluxes can be seen. The verticle dotted
409 lines traversing all three panels indicate the midpoints of the interface
410 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
411
412 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
413 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
414 by eq. 6\cite{Louden13},we find the interfacial width to be
415 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
416 momentum flux for both the pyramidal and secondary prismatic systems.
417 Over the range of shear rates investigated,
418 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
419 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
420 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
421 significant change in the interfacial width. This follows our previous
422 findings of the basal and
423 prismatic systems, in which the interfacial width was invarient of the
424 shear rate of the ice. The interfacial width of the quiescent basal and
425 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
426 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
427 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
428 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
429 $\mathrm{ms}^{-1}$ for the prismatic.
430
431 These results indicate that the surface structure of the exposed ice crystal
432 has little to no effect on how far into the bulk the ice-like structural
433 ordering is. Also, it appears that the interface is not structurally effected
434 by the movement of water over the ice.
435
436
437 \subsection{Orientational dynamics}
438 %Should we include the math here?
439 The orientational time correlation function,
440 \begin{equation}\label{C(t)1}
441 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
442 \end{equation}
443 helps indicate the local environment around the water molecules. The function
444 begins with an initial value of unity, and decays to zero as the water molecule
445 loses memory of its former orientation. Observing the rate at which this decay
446 occurs can provide insight to the mechanism and timescales for the relaxation.
447 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
448 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
449 an ensemble average over all the water molecules in a given spatial region.
450
451 To investigate the dynamics of the water molecules across the interface, the
452 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
453 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
454 molecule was allocated to a particular bin if it was initially in the bin
455 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
456 followed by an additional 200 ps NVE simulation during which the
457 position and orientations of each molecule were recorded every 0.1 ps.
458
459 The data obtained for each bin was then fit to a triexponential decay
460 with the three decay constants
461 $\tau_{short}$ corresponding to the librational motion of the water
462 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
463 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
464 motion of the water molecules. An additive constant in the fit accounts
465 for the water molecules trapped in the ice which do not experience any
466 long-time orientational decay.
467
468 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
469 profiles for the three decay constants, $\tau_{short}$ (panel a),
470 $\tau_{middle}$ (panel b),
471 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
472 respectively. The control experiments (no shear) are shown in black, and
473 an experiment with an imposed momentum flux is shown in red. The vertical
474 dotted line traversing all three panels denotes the midpoint of the
475 interface as determined by the local tetrahedral order parameter fitting.
476 In the liquid regions of both systems, we see that $\tau_{middle}$ and
477 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
478 resepctively, and increase in value as we approach the interface. Conversely,
479 in panel a, we see that $\tau_{short}$ decreases from the liquid value
480 of $72-76$ fs as we approach the interface. We believe this speed up is due to
481 the constrained motion of librations closer to the interface. Both the
482 approximate values for the decays and trends approaching the interface match
483 those reported previously for the basal and prismatic interfaces.
484
485 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
486 and $d_{secondary prismatic}$, from the
487 interface that the deviations from the bulk liquid values begin. This was done
488 by fitting the orientational decay constant $z$-profiles by
489 \begin{equation}\label{tauFit}
490 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
491 \end{equation}
492 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
493 values of the decay constants, $z_{wall}$ is the location of the interface,
494 and $d$ is the displacement from the interface at which these deviations
495 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
496 determined
497 for each of the decay constants, and then averaged for better statistics
498 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
499 system,
500 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
501 system. We found $d_{secondary prismatic}$ to be slightly larger than
502 $d_{pyramidal}$ for both the control and with an applied shear, with
503 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
504 experiment with the imposed momentum flux. These values are consistent with
505 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
506 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
507
508
509
510
511
512 \section{Conclusion}
513 We present the results of molecular dynamics simulations of the basal,
514 prismatic, pyrmaidal
515 and secondary prismatic facets of an SPC/E model of the
516 Ice-I$_\mathrm{h}$/water interface, and show that the differential
517 coefficients of friction among the four facets are due to their
518 relative hydrophilicities by means
519 of water contact angle calculations. To obtain the coefficients of
520 friction, the ice was sheared through the liquid
521 water while being exposed to a thermal gradient to maintain a stable
522 interface by using the minimally perturbing VSS RNEMD method. Water
523 contact angles are obtained by fitting the spreading of a liquid water
524 droplet over the crystal facets.
525
526 In agreement with our previous findings for the basal and prismatic facets, the interfacial
527 width of the prismatic and secondary prismatic crystal faces were
528 found to be independent of shear rate as measured by the local
529 tetrahedral order parameter. This width was found to be
530 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
531 These values are in good agreement with our previously calculated interfacial
532 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
533 systems.
534
535 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
536 by calculation of the orientational time correlation function at varying
537 displacements normal to the interface. The decays were fit
538 to a tri-exponential decay, where the three decay constants correspond to
539 the librational motion of the molecules driven by the restoring forces of
540 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
541 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
542 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
543 $\tau_{short}$ was found to decrease approaching the interface due to the
544 constrained motion of the molecules as the local environment becomes more
545 ice-like. Conversely, the two longer-time decay constants were found to
546 increase at small displacements from the interface. As seen in our previous
547 work on the basal and prismatic facets, there appears to be a dynamic
548 interface width at which deviations from the bulk liquid values occur.
549 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
550 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
551 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
552 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
553 facets, no apparent dependence of the dynamic width on the shear rate was
554 found.
555
556 The interfacial friction coefficient, $\kappa$, was determined for each facet
557 interface. We were able to reach an expression for $\kappa$ as a function of
558 the velocity profile of the system which is scaled by the viscosity of the liquid
559 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
560 independent of temperature differences of the liquid water at far displacements
561 from the interface. We found the basal and pyramidal facets to have
562 similar $\kappa$ values, of $\kappa \approx$ 6
563 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
564 prismatic and secondary prismatic facets were found to have $\kappa$ values of
565 $\kappa \approx$ 3 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
566 Believing this difference was due to the relative hydrophilicities of
567 the crystal faces, we have calculated the infinite decay of the water
568 contact angle, $\theta_{\infty}$, by watching the spreading of a water
569 droplet over the surface of the crystal facets. We have found
570 $\theta_{\infty}$ for the basal and pyramidal faces to be $\approx$ 34
571 degrees, while obtaining $\theta_{\infty}$ of $\approx$ 43 degrees for
572 the prismatic and secondary prismatic faces. This indicates that the
573 basal and pyramidal faces of ice-I$_\mathrm{h}$ are more hydrophilic
574 than the prismatic and secondary prismatic. These results also seem to
575 explain the differential friction coefficients obtained through the
576 shearing simulations, namely, that the coefficients of friction of the
577 ice-I$_\mathrm{h}$ crystal facets are governed by their inherent
578 hydrophilicities.
579
580
581 \begin{acknowledgments}
582 Support for this project was provided by the National
583 Science Foundation under grant CHE-1362211. Computational time was
584 provided by the Center for Research Computing (CRC) at the
585 University of Notre Dame.
586 \end{acknowledgments}
587
588 \bibliography{iceWater}
589 % *****************************************
590 % There is significant interest in the properties of ice/ice and ice/water
591 % interfaces in the geophysics community. Most commonly, the results of shearing
592 % two ice blocks past one
593 % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
594 % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
595 % simulations, Samadashvili has recently shown that when two smooth ice slabs
596 % slide past one another, a stable liquid-like layer develops between
597 % them\cite{Samadashvili13}. To fundamentally understand these processes, a
598 % molecular understanding of the ice/water interfaces is needed.
599
600 % Investigation of the ice/water interface is also crucial in understanding
601 % processes such as nucleation, crystal
602 % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
603 % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
604 % properties can also be applied to biological systems of interest, such as
605 % the behavior of the antifreeze protein found in winter
606 % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
607 % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
608 % give rise to these processes through experimental techniques can be expensive,
609 % complicated, and sometimes infeasible. However, through the use of molecular
610 % dynamics simulations much of the problems of investigating these properties
611 % are alleviated.
612
613 % Understanding ice/water interfaces inherently begins with the isolated
614 % systems. There has been extensive work parameterizing models for liquid water,
615 % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
616 % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
617 % ($\dots$), and more recently, models for simulating
618 % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
619 % melting point of various crystal structures of ice have been calculated for
620 % many of these models
621 % (SPC\cite{Karim90,Abascal07}, SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Gernandez06,Abascal07,Vrbka07}, TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07}, TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}),
622 % and the partial or complete phase diagram for the model has been determined
623 % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
624 % Knowing the behavior and melting point for these models has enabled an initial
625 % investigation of ice/water interfaces.
626
627 % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
628 % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
629 % done significant work characterizing and quantifying the width of these
630 % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
631 % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
632 % recent years, Haymet has focused on investigating the effects cations and
633 % anions have on crystal nucleaion and
634 % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
635 % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
636 % surface restructuring at temperatures approaching the melting
637 % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
638 % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
639 % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
640 % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
641 % this model, Nada and Furukawa have established differential
642 % growth rates for the basal, prismatic, and secondary prismatic facets of
643 % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
644 % bond network in water near the interface\cite{Nada05}. While the work
645 % described so far has mainly focused on bulk water on ice, there is significant
646 % interest in thin films of water on ice surfaces as well.
647
648 % It is well known that the surface of ice exhibits a premelting layer at
649 % temperatures near the melting point, often called a quasi-liquid layer (QLL).
650 % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
651 % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
652 % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
653 % Similarly, Limmer and Chandler have used course grain simulations and
654 % statistical field theory to estimated QLL widths at the same temperature to
655 % be about 3 nm\cite{Limmer14}.
656 % Recently, Sazaki and Furukawa have developed an experimental technique with
657 % sufficient spatial and temporal resolution to visulaize and quantitatively
658 % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
659 % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
660 % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
661 % QLLs, which displayed different stabilities and dynamics on the crystal
662 % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
663 % of the crystal facets would help further our understanding of the properties
664 % and dynamics of the QLLs.
665
666 % Presented here is the follow up to our previous paper\cite{Louden13}, in which
667 % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
668 % investigated where the ice was sheared relative to the liquid. By using a
669 % recently developed velocity shearing and scaling approach to reverse
670 % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
671 % velocity gradients can be applied to the system, which allows for measurment
672 % of friction and thermal transport properties while maintaining a stable
673 % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
674 % correlation functions were used to probe the interfacial response to a shear,
675 % and the resulting solid/liquid kinetic friction coefficients were reported.
676 % In this paper we present the same analysis for the pyramidal and secondary
677 % prismatic facets, and show that the differential interfacial friction
678 % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
679 % relative hydrophilicity by means of dynamics water contact angle
680 % simulations.
681
682 % The local tetrahedral order parameter, $q(z)$, is given by
683 % \begin{equation}
684 % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
685 % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
686 % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
687 % \label{eq:qz}
688 % \end{equation}
689 % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
690 % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
691 % molecules $i$ and $j$ are two of the closest four water molecules
692 % around molecule $k$. All four closest neighbors of molecule $k$ are also
693 % required to reside within the first peak of the pair distribution function
694 % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
695 % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
696 % for the varying population of molecules within each finite-width bin.
697
698
699 % The hydrophobicity or hydrophilicity of a surface can be described by the
700 % extent a droplet of water wets the surface. The contact angle formed between
701 % the solid and the liquid, $\theta$, which relates the free energies of the
702 % three interfaces involved, is given by Young's equation.
703 % \begin{equation}\label{young}
704 % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
705 % \end{equation}
706 % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
707 % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
708 % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
709 % wettability and hydrophobic surfaces, while small contact angles
710 % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
711 % hydrophilic surfaces. Experimentally, measurements of the contact angle
712 % of sessile drops has been used to quantify the extent of wetting on surfaces
713 % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
714 % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
715 % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
716 % Luzar and coworkers have done significant work modeling these transitions on
717 % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
718 % the change in contact angle to be due to the external field perturbing the
719 % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
720
721
722
723 \end{article}
724
725 \begin{figure}
726 \includegraphics[width=\linewidth]{Pyr_comic_strip}
727 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
728 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
729 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
730 Middle panel: the imposed thermal gradient required to maintain a fixed
731 interfacial temperature. Upper panel: the transverse velocity gradient that
732 develops in response to an imposed momentum flux. The vertical dotted lines
733 indicate the locations of the midpoints of the two interfaces.}
734 \end{figure}
735
736 \begin{figure}
737 \includegraphics[width=\linewidth]{SP_comic_strip}
738 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
739 rate of 3.5 \
740 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
741 \end{figure}
742
743 \begin{figure}
744 \includegraphics[width=\linewidth]{Pyr-orient}
745 \caption{\label{fig:PyrOrient} The three decay constants of the
746 orientational time correlation function, $C_2(t)$, for water as a function
747 of distance from the center of the ice slab. The vertical dashed line
748 indicates the edge of the pyramidal ice slab determined by the local order
749 tetrahedral parameter. The control (black circles) and sheared (red squares)
750 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
751 shown by the black and red lines respectively. The upper two panels show that
752 translational and hydrogen bond making and breaking events slow down
753 through the interface while approaching the ice slab. The bottom most panel
754 shows the librational motion of the water molecules speeding up approaching
755 the ice block due to the confined region of space allowed for the molecules
756 to move in.}
757 \end{figure}
758
759 \begin{figure}
760 \includegraphics[width=\linewidth]{SP-orient-less}
761 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
762 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
763 \end{figure}
764
765
766 \begin{table}[h]
767 \centering
768 \caption{Droplet and Shearing simulation parameters}
769 \label{tab:method}
770 \begin{tabular}{|cccc|ccc|} \hline
771 & \multicolumn{3}{c}{Droplet} & \multicolumn{3}{c|}{Shearing}\\
772 Interface & $N_{ice}$ & $N_{droplet}$ & Lx, Ly (\AA) &
773 $N_{ice}$ & $N_{liquid}$ & Lx, Ly, Lz (\AA) \\ \hline
774 Basal & 12960& 2048 & (134.70, 140.04) & 900 & 1846 & (23.87, 35.83, 98.64)\\
775 Prismatic & 9900& 2048 & (110.04, 115.00) & 3000 & 5464 &
776 (35.95, 35.65, 205.77)\\
777 Pyramidal & 11136 & 2048& (143.75, 121.41) & 1216 & 2203 &
778 (143.75, 121.41)\\
779 Secondary Prismatic & 11520 & 2048 & (146.72, 124.48) & 3840 &
780 8176 & (71.87, 31.66, 161.55)\\
781 \hline
782 \end{tabular}
783 \end{table}
784
785
786 \begin{table}[h]
787 \centering
788 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
789 \label{tab:kappa}
790 \begin{tabular}{|ccc|cccc|} \hline
791 & \multicolumn{2}{c}{Droplet} & \multicolumn{4}{c|}{Shearing}\\
792 Interface & $\theta^{\circ}_{\infty}$ & $K_{spread} (ns^{-1})$ &
793 $\kappa_{x}$ & $\kappa_{y}$ & d$_{q_{z}}$ (\AA) & d$_{\tau}$ (\AA) \\ \hline
794 Basal & $34.1 \pm 0.9$ &$0.60 \pm 0.07$ & $5.9 \pm 0.3$&$6.5 \pm 0.8$
795 & $3.2 \pm 0.4$ & $2.9$ \\
796 Pyramidal & $35 \pm 3$ & $0.7 \pm 0.1$ & $5.8 \pm 0.4$ & $6.1 \pm
797 0.5$ & $3.2 \pm 0.2$ & $2.7$ \\
798 Prismatic & $45 \pm 3$ & $0.75 \pm 0.09$ & $3.0 \pm 0.2$ & $3.0 \pm
799 0.1$ & $3.6 \pm 0.2$ & $3.5$ \\
800 Secondary Prismatic & $42 \pm 2$ & $0.69 \pm 0.03$ & $3.5 \pm 0.1$ &
801 $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $3$ \\ \hline
802 \end{tabular}
803 \end{table}
804
805 \end{document}